Download Physics and Computation of Aero-Optics

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Fourier optics wikipedia , lookup

Ellipsometry wikipedia , lookup

Fiber-optic communication wikipedia , lookup

Thomas Young (scientist) wikipedia , lookup

Ultraviolet–visible spectroscopy wikipedia , lookup

Anti-reflective coating wikipedia , lookup

Laser beam profiler wikipedia , lookup

Optical coherence tomography wikipedia , lookup

Optical rogue waves wikipedia , lookup

Nonimaging optics wikipedia , lookup

Silicon photonics wikipedia , lookup

Photon scanning microscopy wikipedia , lookup

Optical aberration wikipedia , lookup

Magnetic circular dichroism wikipedia , lookup

Interferometry wikipedia , lookup

Passive optical network wikipedia , lookup

3D optical data storage wikipedia , lookup

Harold Hopkins (physicist) wikipedia , lookup

Optical tweezers wikipedia , lookup

Nonlinear optics wikipedia , lookup

Transcript
1
Aero-Optics
Physics and Computation of Aero-Optics
Meng Wang
Department of Aerospace and Mechanical Engineering, University of Notre
Dame, Notre Dame, IN 46556; email: [email protected]
Ali Mani
Department of Chemical Engineering, Massachusetts Institute of Technology,
Cambridge, MA 02139; email: [email protected]
Stanislav Gordeyev
Department of Aerospace and Mechanical Engineering, University of Notre
Dame, Notre Dame, IN 46556; email: [email protected]
Key Words wavefront distortion, turbulent flow, high-fidelity simulation, optical mitigation
Abstract
This article provides a critical review of aero-optics with an emphasis on recent developments
in computational predictions and the physical mechanisms of flow-induced optical distortions.
Following a brief introduction of the fundamental theory and key concepts, computational techniques for aberrating flow fields and optical propagation are discussed along with a brief survey
Annu. Rev. Fluid Mech. 2012. 44
of wavefront sensors used in experimental measurements. New physical understanding generated
through numerical and experimental investigations are highlighted for a numbers of important
aero-optical flows including turbulent boundary layers, separated shear layers and flow over
optical turrets. Approaches for mitigating aero-optical effects are briefly discussed.
1
INTRODUCTION
Distortions of optical signals by turbulent flow are widely observed in nature and
in technological applications. The twinkling of stars, or stellar scintillation, is
the result of refraction of light (electro-magnetic waves) by turbulent fluctuations in the Earth’s atmosphere. Shadowgraphs and Schlieren are popular flow
visualization techniques which exploit the optical distortions to visualize the flow
field that produced them. Over the past four decades, much attention has been
paid to the aero-optical effects on the performance of airborne laser systems for
communication, target tracking and directed-energy weapons. In these systems
optical distortions produced by turbulent flows surrounding the projection aperture pose a serious problem, causing beam distortion, jitter and much reduced
effective range. The performance of airborne and ground-based imaging systems
is likewise impaired by turbulent flows in the vicinity of the viewing aperture.
The direct cause of optical distortions is the density variations in the flow field.
For air and many other fluids, the index of refraction is linearly related to the
density of the fluid by the Gladstone-Dale relation (Wolfe & Zizzis 1978). When
a beam of an initially planar wavefront is transmitted through a variable density
field, different parts of the beam propagate at different local speed of light, resulting in distortions of the wavefront. An optical beam emitted from a projection
aperture, or received by a viewing aperture, typically transmits/receives through
two distinct flow regions: the active turbulence region induced by solid objects
2
Aero-Optics
3
near the optical window, and atmospheric turbulence. The propagation through
atmospheric turbulence has been studied extensively (Chernov 1960, Tatarski
1961) and is relatively well understood. Because of the large temporal and spatial
scales associated with atmospheric turbulence, its aberrating effects are of low
frequency (< 100 Hz) and can be largely corrected using Adaptive-Optic (AO)
systems (e.g. Lloyd-Hart 2003, Hardy 1998). In contrast, the turbulent flow induced by solid surfaces near the aperture, which may be comprised of turbulent
boundary layers, free shear-layers, wakes, and shock waves for supersonic and
transonic flows, is characterized by much smaller turbulence scales. The size of
the optically-active flow region is typically thinner than or comparable to the
aperture size. Aero-optics, to follow the conventional definition (Gilbert & Otten
1982, Sutton 1985, Jumper & Fitzgerald 2001), is concerned with the aberrating
effects of compressible turbulence in this region. Compared to the atmospheric
boundary layer, the aero-optical flow generates stronger optical aberrations at
smaller scales and higher frequencies, which are beyond the capability of today’s
AO technology. Mitigation of these aberrations via active and passive flow-control
has been actively pursued in recent years (e.g. Gordeyev et al. 2010a, 2010b).
The types of aero-optical distortions and their far-field impact depend on a
number of physical and geometric parameters, including the optical wavelength
λ, aperture or beam size a, turbulence length scale ℓ of the aberrating field, and
distance of propagation L. In crude terms, small-scale turbulence, with eddy sizes
less than the optical aperture (ℓ < a), causes optical scattering, beam spread,
and consequent attenuation of intensity, whereas turbulent eddies larger than
or comparable in size to the aperture (ℓ ≥ a) are mostly responsible for the unsteady tilt of the beam (beam jitter) (Cassady et al. 1989). In addition, the mean
4
WANG, MANI & GORDEYEV
density gradient in the beam path causes steady wavefront distortions known as
lensing effect. The magnitude of wavefront distortions OPDrms (its precise definition will be given later) is generally small in an absolute sense but can be a
significant fraction of, or even exceed the optical wavelength. In other words,
the optical phase distortion 2πOPDrms /λ, which determines the far-field beam
quality, can be easily of O(0.1) to O(1). For the same distortion magnitude, the
phase distortion of an aberrated beam is inversely proportional to the optical
wavelength, making the aero-optics problem particularly acute for short wavelength beams (Jumper & Fitzgerald 2001). This poses a significant impediment
to using shorter wavelength laser systems, which are preferable in the absence of
aero-optical aberrations; for an unaberrated beam the diffraction-limited optical
intensity scales with a2 /(L2 λ2 ) (Born & Wolf 2002).
A number of review articles have been written on aero-optics in the past. Research in the pre-1980 era is documented in Gilbert & Otten (1982). Progress
since then has been surveyed by Sutton (1985), Jumper & Fitzgerald (2001),
and more recently Gordeyev & Jumper (2010) with a focus on the aero-optics of
turrets. Among recent advances in this field are high-speed and high-resolution
wavefront sensors which allow measurements of instantaneous wavefront errors
in unprecedented detail, and the increasing role played by numerical simulations. Application of computational fluid dynamics (CFD) to aero-optics problems has historically lagged most other areas because of the challenging nature
of the computations, which must be time-accurate, compressible, and capture
optically important flow scales. However, this has begun to change with recent
advances in high-fidelity simulation tools and the concomitant increase in computing power. A new generation of computational techniques including direct
Aero-Optics
5
numerical simulation (DNS), large-eddy simulation (LES), and hybrid methods
combining LES with Reynolds-averaged Navier-Stokes (RANS) approaches, have
been employed hand-in-hand with experimental and theoretical approaches to
elucidate the physics of aero-optics, predict aberration effects, and develop techniques for their mitigation. This article is an attempt to provide a critical review
of some recent progress in the understanding and prediction of aero-optical effects. The discussion is primarily from a computational perspective, although
important experimental findings are also included.
2
2.1
THEORETICAL BACKGROUND
Basic Equations
The theoretical foundation for electromagnetic wave propagation in a turbulent
medium is discussed extensively in Monin & Yaglom (1975), and is briefly outlined
here in the context of aero-optics. In the most general sense, the propagation of
electromagnetic waves is governed by the Maxwell equations along with compressible Navier-Stokes equations. Various simplifications can be made depending on
relevant physical parameters and length and time scales. For aero-optical problems, the time scale for optical propagation is negligibly short relative to flow
time scales, and hence optical propagation can be solved in a frozen flow field at
each time instant. If the optical wavelength is much shorter than the smallest
flow scale (Kolmogorov scale), which is generally the case, the effect of depolarization is negligible, and the Maxwell equations reduce to a vector wave equation
in which all three components of the electromagnetic field are decoupled. A scalar
6
WANG, MANI & GORDEYEV
component of the electric field at frequency ω is governed by
∇2 U +
ω 2 n2
U = 0,
c20
(1)
where c0 is the speed of light in vacuum and n is the index of refraction. The
latter is related to the density of air via the Gladstone-Dale relation: n(x, y, z) =
1+KGD (λ)ρ(x, y, z), where KGD is the Gladstone-Dale constant and is in general
weakly dependent on the optical wavelength (Wolfe & Zizzis 1978). In aerooptics, fluctuations in index of refraction are small (∼ 10−4 ) and have scales
much larger than the optical wavelength, and hence an optical beam propagates
predominantly in the axial (z) direction with slowly varying amplitude. It is
customary to invoke the paraxial approximation, which assumes U (x, y, z) =
A(x, y, z) exp(−ikz) and |∂ 2 A/∂z 2 | ≪ |k∂A/∂z|, where k = ω/c∞ is the optical
wavenumber in the free stream. This leads to the parabolized wave equation for
the complex amplitude
∂A
−2ik
+ ∇2⊥ A + k2
∂z
n2
− 1 A = 0,
n2∞
(2)
where ∇2⊥ is the Laplacian operator in the transverse directions. Effects not included in the paraxial wave equation include large-angle scattering and energy
dissipation/absorption, which are insignificant for aero-optical problems. A detailed discussion of the approximations and limitation of applicability is given by
Monin & Yaglom (1975).
Further approximations can be made to obtain closed-form solutions to Equation 2. As depicted schematically in Figure 1, the propagation domain consists
of two distinct regions: the optically active turbulent near-field (from z = 0 to
z1 ) and the free space extending to the distant far field (from z = z1 to L, with
L ≫ z1 ). For transmission across the aero-optical region, it can be shown that the
7
Aero-Optics
contribution from the Laplacian term in Equation 2, which represents the diffraction effect, is negligible relative to the last term for z1 ≪ ℓ2⊥ /λ, where ℓ⊥ is the
transverse length scale of the optical wave. If the refractive index is expressed as
n = n∞ +n′ , and the linear approximation n2 /n2∞ ≈ 1+2n′ /n∞ is made about the
index-of-refraction deviation from the free-stream value, Equation 2 can be inte
Rz
grated from 0 to z1 to obtain A(x, y, z1 ) = A(x, y, 0) exp − nik∞ 0 1 n′ (x, y, z)dz ,
or
U (x, y, z1 ) = U (x, y, 0) exp −ik0
Z
z1
n(x, y, z)dz ,
0
(3)
where k0 = k/n∞ = ω/c0 = 2π/λ0 is the optical wavenumber in vacuum.
Once the solution past the aero-optical region, U (x, y, z1 ), has been determined,
it can be used as the initial condition to propagate the optical beam to the far field
using the wave equation (Equation 1) or paraxial wave equation (Equation 2).
By setting n = n∞ and n′ = 0, either equation can be solved analytically using
Fourier transform techniques, which are the basis for Fourier optics (Goodman
2004).
2.2
Near-Field Distortion Measures
Equation 3 suggests that the dominant aero-optical effect after transmission
through the turbulence region is a phase distortion of the optical wavefront;
the amplitude is approximately unchanged. The integral in Equation 3 is known
as the optical path length (OPL). It is most commonly derived from geometric
optics by assuming straight ray paths and is generally dependent on the flow-time
scale (but not optical time scale):
OPL(x, y, t) =
Z
z1
n(x, y, z, t)dz.
0
(4)
8
WANG, MANI & GORDEYEV
Ray bending can be accounted for by solving the eikonal equation (Born & Wolf
2002) but is generally negligible for aero-optics. In practice, the relative difference
in the OPL over the aperture is a more relevant representation of wavefront
distortions. It is called the optical path difference (OPD) and is defined as
OPD(x, y, t) = OPL(x, y, t) − hOPL(x, y, t)i,
(5)
where the angle brackets denote spatial averaging over the aperture. The optical
phase distortion is then 2πOPD/λ. It should be noted that the optical wavefront,
defined as the locus of constant phase, is (after removing the spatial mean) the
conjugate (negative) of the OPD: W (x, y, t) = −OPD(x, y, t).
To facilitate analysis and mitigation of distortions, the time-dependent OPD
is often decomposed into a time-averaged spatial component, called the steadylensing term, OPDsteady (x, y), and an unsteady component. The unsteady part
can be further split into a spatially linear component, called unsteady tilt or beam
jitter, and the rest, usually called high-order distortions (Gordeyev & Jumper
2010). In other words,
OPD(x, y, t) = OPDsteady (x, y) + [A(t)x + B(t)y] + OPDhigh−order (x, y, t). (6)
Physically, these three components affect an outgoing beam in different ways. The
steady-lensing term, OPDsteady (x, y), is a function of the time-averaged density
field only and imposes a steady distortion like defocus, coma and so on. The tilt
or jitter, represented by the second term on the right-hand side, does not change
the spatial distribution of the outgoing beam but simply re-directs it in directions
defined by functions A(t) and B(t). The specific forms of A(t) and B(t) depend
on the definition of tilt. For the so-called G-tilt (Tyson 2000), A(t) and B(t)
are the spatially-average gradient components of the OPD in x- and y-directions,
9
Aero-Optics
respectively, whereas for the Z-tilt (Sasiela 2007), they are defined such that the
magnitude of OPDhigh−order in Equation 6 is minimized in the least square sense
at each time instant. Finally, the high-order term causes the beam to change its
shape and intensity distribution.
The decomposition in Equation 6 is particularly useful when an adaptive optic
system is used to correct for aberrating wavefronts. The purpose of an adaptive
optic system is to apply a conjugate wavefront to the outgoing beam so that
it will negate optical aberrations from the flow and the beam will become recollimated after passing through the turbulent media (Tyson 1997). Among the
distortion components given in Equation 6, steady lensing is easily corrected by
a deformable mirror with a large range of motion, the tilt component is removed
using a fast-steering mirror, and the high-order term can be compensated for by
using a high-bandwidth deformable mirror (Tyson 1997).
2.3
Statistical Theory
An important equation relating the statistical properties of a turbulent medium
and those of aero-optical aberrations, the so-called linking equation, was derived
by Sutton (1969) (see also Sutton 1985, Steinmetz 1982, and Havener 1992). In
the most general form, the linking equation can be written as,
2
hOPD i =
2
KGD
Z
0
z1
Z
z1
Covρ′ (z, z ′ )dz ′ dz,
(7)
0
where the overbar denotes time averaging, Covρ′ is the covariance function of
density fluctuations, and z1 is the integration distance along the traversing beam
through the turbulence region. In the case of homogeneous turbulent flows,
the density covariance is most commonly modeled by an exponential function
(Steinmetz 1982), ρ2rms exp(− |z − z ′ | /Λ), or a Gaussian function (Wolters 1973),
10
WANG, MANI & GORDEYEV
ρ2rms exp(− |z − z ′ |2 /Λ2 ), where Λ is the characteristic length scale for density
fluctuations. Substituting these models into Equation 7 leads to
2
hOPD i =
2
αKGD
Z
z1
ρ2rms (z)Λ(z)dz,
(8)
0
where α = 2 for the exponential covariance function and
√
π for the Gaussian
covariance function. Both the density fluctuation magnitude and length scale are
allowed to vary slowly along the beam path.
Both forms of the linking equation allow one to calculate aero-optical distortions indirectly from statistical properties of the turbulent flow. Since the full
covariance matrix in Equation 7, is difficult to measure experimentally, the simplified equation, Equation 8 is commonly used instead. However, since the simplified linking equation is derived for homogenous and isotropic turbulent flows,
its applicability to inhomogeneous flows, such as shear layers (Hugo & Jumper
2000) and boundary layers (Gilbert & Otten 1982, Tromeur et al. 2006b), has
been questioned. It has been shown that with appropriate choice of the length
scale Λρ , the simplified linking equation can be used to obtain accurate results for
flow fields with anisotropic and inhomogeneous turbulence. The key is to use the
correct density correlation length defined based on Equation 7 (Wang & Wang
2011).
3
PREDICTION OF FAR-FIELD DISTORTIONS
Given the OPD profiles after the beam passes the turbulence region, its free-space
propagation from z = z1 to L can be solved using Fourier optics to obtain the
exact far-field projection. Figure 2 shows an example of instantaneous far-field
irradiance of a Gaussian beam subject to strong aero-optical distortions by the
turbulent wake behind a circular cylinder of diameter D (Mani et al. 2009). It
Aero-Optics
11
contrasts the distorted beam pattern to that of an undistorted, diffraction limited
beam for two different optical wavelengths. In this case the near-field OPDrms
is 6.7 × 10−6 D, corresponding to phase distortions of 2πOPDrms /λ ≈ 17 and
4.2 for the two wavelengths (6.7 and 1.7 respectively, after tilt removal). Drastic
losses of far-field beam intensity and coherence are observed. Note that in the
absence of distortions, the shorter wavelength beam delivers more energy to the
target than the longer wavelength one, but this advantage is offset by aero-optical
distortions. In fact, the ratio of the maximum intensity of the distorted beam
to that of the undistorted one is about 1.5% for the shorter wavelength case and
10% for the longer wavelength case.
While Fourier optics provides a complete description of the optical propagation,
it requires as input detailed spatio-temporal history of the near-field wavefront,
which may not available from experimental measurements. Furthermore, from a
systems point of view, statistical measures, rather than instantaneous quantities,
are of primary interest, and it is desirable to relate the far-field beam statistics to
statistics of the aberrating flow field. A widely-adopted relation is the Maréchal
approximation (Maréchal 1947, Born & Wolf 2002) for the Strehl ratio (SR),
defined as the ratio of the peak on-axis far-field irradiance of an aberrated beam
(after tilt removal) to the corresponding peak irradiance of an unaberrated beam;
it is a measure of beam quality relative to a diffraction-limited beam at each time
instant. In his original paper Maréchal (1947) showed that, in the limit of small
phase distortions, OPDrms /λ < 0.1, SR(t) = I(t)/I0 ≈ 1 − [2πOPDrms (t)/λ]2 . A
more commonly adopted expression, however, takes the exponential form
" #
I(t)
2πOPDrms (t) 2
.
= SR(t) ≈ exp −
I0
λ
(9)
This equation, also known as Maréchal approximation, was proposed by Mahajan
12
WANG, MANI & GORDEYEV
(1982, 1983) based on empirical fitting of SR versus OPDrms data for a variety
of distortion modes. In comparison with the original form, Mahajan found that
Equation 9 gave a better approximation for the Strehl ratio over a wider range
of phase distortions.
Ross (2009) recently re-examined the derivation of the Maréchal approximation
and found that for OPD data with Gaussian probability-density distribution,
Equation 9 was in fact exact. It is important to note that the original Maréchal
approximation is a relationship between the spatial statistics of the instantaneous
wavefront and the far-field intensity. In the case when the aperture size is much
larger than the turbulence length scale, a ≫ l, Steinmetz (1982) showed that the
Maréchal approximation is also approximately valid for the time-averaged Strehl
ratio:
" #
2πOPDrms 2
.
SR ≈ exp −
λ
(10)
This equation is known interchangeably as the large-aperture approximation and
Maréchal approximation. However, in many practical applications, the assumptions of large (infinite) aperture and small aberrations are not exactly valid.
For stationary, Gaussian-in-space processes, an exact version of time-averaged
Maréchal approximation can be expressed in terms of the temporal probabilitydensity function of OPDrms (Porter et al. 2011).
The time-averaged version of Maréchal approximation is currently the de facto
standard approach for estimating and comparing the effect of aero-optical distortions on the degradation of beam performance. Despite its routine use, some
uncertainties still exist concerning the accuracy and applicability of this approximation, particularly for cases with large phase distortions.
Mani et al. (2006) proposed a new set of statistical measures to quantify the
Aero-Optics
13
far-field optical distortions, and derived exact algebraic relations between these
measures and the statistics of near-field OPD. Instead of characterizing beam
distortions in terms of peak irradiance, they proposed to use the spatial spread of
the beam, in terms of the second spatial moment of the far-field irradiance, as a
measure of distortion (see Figure 2). Through statistical solutions of the paraxial
wave equation, it was shown that the square of the beam spread about its center
(i.e. with tilt removal) can be expressed, in the Fraunhofer limit Lλ/a2 ≫ 1, as
*
′2 +
σ 2
λ2
∂
x
.
(11)
=C 2 2 +
OPD
L
π a
∂x
where σx is the beam spread in the x-direction, a is the aperture diameter, the
prime represents deviation from the mean value, and the angle brackets indicate
a spatial average. C is a constant determined by the shape of the aperture and its
optical intensity profile; it is an O(1) quantity for beams with smooth aperture
intensity profiles and is exactly equal to one for Gaussian beams. The beam
spread in the y-direction (see Figure 2) has the same form. One can observe
clearly that the anisotropy of the beam spread about its axis is directly linked to
the anisotropy of OPD gradients and thus in the refractive index field.
The first term on the right-hand side of Equation 11 represents the effect of
diffraction on the beam spread, which is decoupled from the aero-optical effect
represented by the second term. The latter is simply the variance of OPD gradient. The ratio of these two terms was proposed by Mani et al. (2006) as the basis
for a new distortion measure, which they termed the “fidelity ratio”. Compared
with the Strehl ratio based on Maréchal approximation, the new measure has the
advantage of being applicable to highly aberrated beams because Equation 11 is
exact. The disadvantage is that its applicability is limited to beams with initially
smooth (continuous) aperture intensity profiles since the beam spread becomes a
14
WANG, MANI & GORDEYEV
singular measure for beams with discontinuous intensity profile. Equation 11 also
provides insight into scaling of aero-optical distortions with key optical parameters such as the aperture size. For instance, one can expect that the aero-optical
effect is a lesser concern for naked human eyes (with a ∼ 1 mm) compared to
tactical airborne system (with a ∼ 20 cm), since the far-field resolution of an eye
is predominantly limited by the diffraction term for typical aero-optical flows.
4
COMPUTATIONAL AND EXPERIMENTAL APPROACHES
4.1
4.1.1
Computational Methods
Optical Propagation
Computations of aero-optics consist of two
essential parts: solutions of the aberrating flow field via CFD techniques and
propagation of the optical beam through the aberrating flow to the target. In
practices, the beam propagation is easily computed by a combination of ray
tracing with Fourier optics. For propagation through the thin aero-optical region
surrounding an aperture, ray optics is applicable, and Equation 3 or its OPD
representation provides satisfactory results. The numerical evaluation involves
a simple integration of the index-of-refraction field along the optical path. The
validity of this approach is established based on scaling arguments, and has been
confirmed numerically by White (2010) through a comparison with the numerical
solution of paraxial wave equation (Equation 2) using a high-order scheme in
the case of optical propagation through a supersonic turbulent boundary layer.
Beyond the aero-optical region, the index of refraction is considered uniform,
and propagation in free space can be treated with Fourier optics as exemplified
in Section 3. If the target is in the optical far field, the Fraunhofer approximation
(Goodman 2004) can be applied to simplify the solution.
15
Aero-Optics
The long-range propagation of optical waves in atmospheric turbulence is significantly more difficult to compute. Ray tracing is invalid because amplitude
variations accumulated over a large distance are significant, and numerical solutions of the wave equation or paraxial wave equation are impeded by the vast
scale disparity and prohibitive computational expenses required to obtain the
fluctuating index-of-refraction field over the entire propagation range. Current
approaches are predominantly based on statistical modeling of the turbulent media in terms discrete thin “phase screens” along the direction of propagation
and numerical integration of the paraxial wave equation (e.g. Coles et al. 1995,
Frehlich 2000). Because its distortion effect can be corrected by adaptive optic
techniques, atmospheric propagation is not considered a pressing issue in aerooptics and is therefore not discussed further.
4.1.2
Computation of Aberrating Flows
The most challenging part
of computational aero-optics is the computation of the aberrating flows surrounding optical apertures. These flows are compressible, turbulent, generally threedimensional at high Reynolds numbers and often with separation. In order to
accurately compute the index of refraction field, turbulence scales over all optically relevant wavenumbers and frequencies must be captured, which poses a
significant challenge in terms of computational expenses and numerical accuracy.
Because of these difficulties, application of CFD to aero-optics did not start until
late 1980s, and significant growth occurred only in the last decade as a result of
the advancement and maturation of high-fidelity simulation techniques.
Limited by computing power, early numerical investigations involved twodimensional solutions of Euler equations and RANS equations. Cassady et al.
(1989) performed RANS simulations of the flow over an open cavity in the Mach
16
WANG, MANI & GORDEYEV
number range between 0.6 and 0.8 under standard conditions at altitudes of
40,000 to 50,000 feet, and analyzed image distortions at different look angles
across the shear layer. Because their RANS calculations did not capture the
coherent structures in the shear layer, only the steady tilt of the optical beam
was predicted directly. Beam jitter was modeled, and high-order distortions were
largely unaccounted for. Tsai and Christiansen (1990) solved Euler equations
numerically to study the degradation of a laser beam by a plane mixing layer
at low Mach numbers (M1 = 0.2 and M2 = 0.1 for the two air streams). Their
solutions were able to capture qualitative features of large-scale coherent structures including vortex pairing, but this is a result of the numerical viscosity in the
simulation code based on a combination of MacCormack and Godunov methods.
While the Euler and RANS solutions are computationally efficient, their usefulness for aero-optical predictions is limited. Euler equations do not describe
the correct physics of refractive-index fluctuations in a turbulent flow and must
rely on numerical dissipation to mimic the effect of physical viscosity. In a steady
RANS calculation, all turbulence scales are modeled, resulting in an ensembleaveraged (time-averaged) density field from which the steady-lensing effect of
optical aberrations can be evaluated but not the unsteady tilt and high-order
effects. It is, however, possible to combine a RANS simulation with a statistical model for optics, such as the linking equation, Equation 8, to estimate the
aero-optical effects. This provides a practical approach for high-Reynolds-number
flows in realistic configurations which is not currently affordable with more accurate techniques. Smith et al. (1990) proposed a transport equation for the
variance of index-of-rafraction fluctuations, and solved it along with compressible
turbulent boundary-layer equations with k-ε model for a plane mixing layer. The
Aero-Optics
17
computed refractive-index variance, together with the turbulence length scale
estimated based on k and ε solutions, allowed the linking equation to predict
the wavefront error for a beam passing the mixing layer and its impact on target intensity. Using the same methodology, Pond & Sutton (2006) performed
an aero-optic analysis of a nose-mounted optical turret on an aircraft. Their
solutions are based on three-dimensional RANS equations, the k-ε model, and
the transport equation for the refractive-index variance proposed by Smith et al.
(1990).
For more accurate computations of index-of-refraction fluctuations, a hierarchy
of high-fidelity methods including DNS, LES and hybrid RANS/LES approaches,
are available. Among them DNS (Moin & Mahesh 1998), which resolves all flow
scales down to the dissipative scale, is the most accurate. Due to the well-known
Reynolds number limitation it is primarily a research tool and has only found
limited use in aero-optics. Truman & Lee (1990) and Truman (1992) used DNS
data to investigate the phase distortions in optical beams through a homogeneous
shear flow with uniform mean shear and a turbulent channel flow. Their simulations were based on incompressible Navier-Stokes equations, and the fluctuating
index-of-refraction field was represented as a passive scalar. The magnitude of
phase distortions was found to be sensitive to the direction of propagation, which
was explained in terms of the anisotropic vortical structures in the flow. Despite
the low Reynolds numbers and incompressible nature of the simulations, the
findings in these studies are qualitatively consistent with subsequent experimental measurements (Cress et al. 2008) and results of compressible flow simulations
(Wang & Wang 2009); see Section 5.1.
LES offers a less expensive alternative to DNS because it resolves only the
18
WANG, MANI & GORDEYEV
large, energy-containing scales of fluid motions which are optically important.
The effect of small scales are modeled by a subgrid-scale (SGS) model. The
prevalent SGS models in use today are of the Smagorinsky eddy-viscosity type
(Smagorinsky 1963) for which the dynamic procedure (Germano et al. 1991,Lilly
1992, Moin et al. 1991) provides a robust way of computing the model coefficient from the resolved scales, thus eliminating the need for adjustable model
parameters. Some practioners use implicit LES, or ILES, in which a high-order
low-pass spatial filter acts as the SGS model to selectively provide numerical
dissipation (e.g. Visbal 2009, White et al. 2010), often out of numerical-stability
consideration for high-order accurate compressible LES codes.
The first application of LES to aero-optics was made by Childs (1993), who
carried out LES of high-speed turbulent mixing layers at convective Mach numbers of up to 2.5 and analyzed the induced wavefront distortions. Significant
growth in LES-based investigations occurred in the past decade as a result of
maturation of LES methodology, advances in computing power and a renewed
interest in the field of aero-optics. LES has to date been employed in a variety
of aero-optical configurations including turbulent shear layers (Visbal 2008, Visbal 2009, White et al. 2010), open-cavity flows (Sinha et al. 2004, Visbal 2008),
turbulent boundary layers (Tromeur et al. 2003, 2006a, 2006b; Wang & Wang
2009, 2011; White & Visbal 2010), flow over a cylinder (Mani et al. 2009), and
flows over cylindrical turrets (Wang & Wang 2009, Morgan & Visbal 2010, Wang
et al. 2010) and realistic three-dimensional turrets (Jones & Bender 2001,Arunajatesan & Sinha 2005). In addition, there have been attempts to employ LES to
predict the effect of passive flow control on aero-optical mitigation (Morgan &
Visbal 2010, Wang et al. 2010). Among the afore-mentioned studies, the simple
Aero-Optics
19
canonical flows such as turbulent shear layers, boundary layers and flow over a
cylinder were simulated to study the fundamental physics of aero-optics and were
afforded adequate grid resolution and numerical accuracy. Some of the findings
from these studies will be discussed in Section 5. On the other hand, the simulations of three-dimensional turret flows were under-resolved and used simulation
codes with large numerical dissipation.
To further reduce computational costs for wall-bounded flows at high Reynolds
numbers, LES can be combined with a RANS model to form a hybrid RANS/LES
method. A popular class of hybrid methods, proposed and recently reviewed by
Spalart (2009), is detached-eddy simulation (DES). In this method less expensive
RANS equations are solved with an appropriate turbulence model in an attached
boundary layer, and LES is used to treat the separated-flow region. It therefore avoids the stringent boundary-layer resolution demanded for LES but at the
same time inherits many weaknesses of RANS methods such as poor prediction
of pressure-driven incipient separation. In addition, the lack of turbulent fluctuations prior to separation may affect the instability of the separated shear layer
and its aero-optical behavior. DES and similar hybrid methods have been applied
in simulations of flow over three-dimensional optical turrets, with limited success
(Nahrstedt et al. 2008, Ladd et al. 2009, Morgan & Visbal 2010); more discussion
will follow in Section 5.3. Another class of hybrid methods, which is more accurate and robust but computationally more demanding than DES, is LES with
wall modeling (Piomelli & Balaras 2002, Wang & Moin 2002). In this case only
the near-wall region in a turbulent boundary layer is treated with RANS modeling, whereas the bulk part of the boundary layer is computed using LES. This
avoids the strong Reynolds number scaling of grid-resolution requirements in a
20
WANG, MANI & GORDEYEV
fully resolved LES. LES with wall modeling has to date not been used for aerooptical computations but holds a bright promise as computing power continues
to grow.
4.2
Resolution Requirement
Turbulent flows relevant to aero-optics contain eddies over a wide range of scales,
typically from order meter to order 10 microns. When practical computational
tools such as LES is employed for aero-optical computations it is important to
ensure that the optical effects of the unresolved flow is negligible. Addressing
this issue requires an understanding of the range of optically active flow scales
in aero-optical flows. Such knowledge not only helps select the mesh size for
computations, but also provides insight for adaptive optics requirements. In
other words, if distortion effects are to be canceled by deformable mirrors, the
range of length scales (and frequencies) to be corrected can be determined from
the same theory.
From Equation 3 it is clear that lack of flow resolution will cause errors in
computation of the optical phase when the beam is traced through the turbulence.
If the true refractive-index field is written in terms of the resolved field, nr , and
the error, ne which is basically the unresolved n, substituting n = nr + ne into
Equation 3 yields:
Z
U (x, y, z1 ) ≃ Ur (x, y, z1 ) − ik0
0
z1
ne (x, y, z)dz Ur (x, y, z1 ),
(12)
where a leading order expansion for small ne is used and Ur is the computed
optical wave (by substituting the resolved field, nr , into Equation 3). According
to this description, the true optical wave at z1 is written in terms of two beams:
the first term on the right hand side represents the computed beam and the
21
Aero-Optics
second term represents the “error” beam associated with lack of resolution in
predicting the refractive index field. An acceptable resolution should ensure that
the energy of the error beam is much smaller than the energy of the resolved
beam.
Mani et al. (2008) used this criterion in conjunction with the Kolmogorov hypothesis for unresolved turbulence to develop a theory to estimate the smallest
optically-important flow length scale in a general aero-optical framework. According to their analysis, in the limit of high Reynolds number the smallest optically important flow scale does not depend on the Kolmogorov scale. For a given
geometry, this length scale depends only on the flow Mach number, freestream
refractive index n∞ , and the optical wavelength λ0 :
6/7
ℓc ≈ C
λ0 ℓ4/7
3/7
M 12/7 (n∞ − 1)6/7 z1
,
(13)
where ℓc is the smallest optically active scale, M is the turbulent Mach number,
z1 is the depth of the turbulent field in the propagation direction, and ℓ is the
length scale of the largest turbulent eddy. The proportionality constant C in
Equation 13 can be written in terms of the universal constant of the Kolmogorov
spectrum and the acceptable energy treshold for the “error” beam in Equation 12,
and is generally of small value; for example, this constant is approximately 0.05
for 95% accuracy (see Mani et al. 2008 for details). A crude estimate based
on practical aero-optical parameters indicates that the length scale predicted by
this theory is in the inertial range and of order of typical LES resolutions. This
indicates that LES, without the need for subgrid optical modeling, is adequate for
aero-optical computations, and DNS is not necessary. Extension of this analysis
to complex flow regimes with inhomogeneous turbulence and mesh spacing can
be found in Mani et al. (2008).
22
WANG, MANI & GORDEYEV
The results of this analysis and the conclusion regarding negligible optical ef-
fects by small-scale turbulence are consistent with previous experimental findings
of Zubair & Catrakis (2007), who verified the “resolution robustness” of optical
distortions by systematically examining the refractive-index data of separated
shear layers at various resolution levels.
4.3
Wavefront Sensors
The most critical device for experimental aero-optics is the wavefront sensor.
The wavefront is the locus of constant phase for the complex intensity (see Equation 3). This quantity cannot be measured directly, as any light-recording device
is sensitive to the absolute value of the complex intensity only. Wavefront sensors
can be divided into several categories depending on how the wavefront is measured, which also defines the sensor performance in terms of the dynamic range,
temporal and spatial resolution and sensitivity to real-life corrupting effects such
as the quality of the optical elements involved or large jitter of the incoming
beam due to mechanical vibration. Interferometry-based sensors combine the
measured wavefront with a known wavefront, which creates an interference pattern with variable light intensity on the sensor, and the measured wavefront can
be directly reconstructed from it. They have very good spatial (tens of thousands of points) and temporal (up to 100 kHz) resolution, but usually require
very good optical elements to manipulate the incoming beam before sending it
to the sensor. Slope-based sensors, like a Shack-Hartmann sensor (Tyson 1997),
rely on the Huygens’ Principle, which states that the wavefront propagates in the
direction normal to itself (Born & Wolf 2002). A lenslet array is used to break the
wavefront into a large number of small sub-apertures and the overall tilt or the
Aero-Optics
23
deflection angle over each sub-aperture is measured by tracing the dot motion at
the lenslet focal plane. The deflection angles are gradients of the wavefront, and
different techniques are used to reconstruct the original wavefront. These sensors
are perhaps the most robust and widely used wavefront sensors with a large dynamic range and relative insensitivity to the quality of the optical components
and the overall beam jitter. The spatial resolution is, however, limited by the
lenselet array, which typically has several thousand sub-apertures, and temporal
resolution is usually limited to several kHz; however, more recent cameras have
now increased this limit to tens of kHz. Finally, phase-diversity and distortedgrating sensors use the intensity transport equation to measure the wavefront
(Blanchard et al. 2000) and ultimately are sensitive to the wavefront curvature.
They have good spatial resolution (several thousand points) and very high temporal resolution (up to 100 kHz), but typically require good optical components to
manipulate the beam, as well as small overall beam jitter, and selective aperture
geometries.
Most of the wavefront devices use high-speed digital cameras to record images, which contain information about measured wavefronts. For slope-based
wavefront sensors, these digital devices limit the sampling speed of data collection. To overcome this problem, analog-only wavefront devices were developed,
which use photo-sensitive diodes to record instantaneous deflection angles over
sub-apertures. Examples of these devices include a Malley Probe (Gordeyev
et al. 2007a), a Small-Aperture Beam Technique (SABT) sensor (Hugo & Jumper
1996) and an analog-only Shack-Hartmann wavefront sensor (Abado et al. 2010).
These devices have high sampling rates (∼ 100 kHz), but either measure only
one-dimensional slice of the wavefront (the Malley probe) or have limited spa-
24
WANG, MANI & GORDEYEV
tial resolution (SABT sensor, analog-only Shack-Hartmann sensor. As a final
note, Equation 13 provides a practical estimate on the minimum number of subapertures needed to correctly measure wavefronts, and for most subsonic flows,
only a few thousand sub-apertures are sufficient for accurate measurements of
OPDrms .
5
AERO-OPTICAL FLOWS AND DISTORTION MECHANISMS
5.1
Turbulent Boundary Layers
The optically-aberrating effects of high-speed, turbulent boundary layers have
been the subject of research since the early 1950s. The first investigation was
by Liepmann (1952) and made use of the jitter angle of a thin beam of light as
it traveled through the compressible boundary layer on the sides of high-speed
wind tunnels as a way to quantify the crispness on Schlieren photographs. Stine
& Winovich (1956) performed photometric measurements of the time-averaged
radiation field at the focal plane of a receiving telescope, and this work also
raised the prospect of using an optical degradation measurement as a method
of inferring turbulence scales. Rose (1979) used hot-wire measurements and the
linking equation (Equation 8) to estimate aero-optical aberrations caused by turbulent boundary layers, and found empirically that OPDrms ∼ qδ, where q is
the dynamic pressure and δ is the boundary-layer thickness. As reviewed by
Gilbert & Otten (1982), research up until 1980 focused on the measurement of
time-averaged optical distortions, either directly by optically-based methods or
indirectly using fluid-mechanical measurements and the linking equation.
The development of high-temporal-resolution Malley probe and other highspeed wavefront devices in recent years allowed accurate time-space-resolved op-
25
Aero-Optics
tical measurements in turbulent boundary layers. Wyckham & Smits (2009)
investigated the aero-optical performance of transonic and supersonic boundary
layers using a two-dimensional Shack-Hartman wavefront sensor. Based on the
OPD equation (Equation 5), and by assuming negligible pressure fluctuations in
the boundary layer and invoking the strong Reynolds analogy (SRA) (Morkovin
1962), they proposed the scaling relation OPDrms ∼ ρM 2 δ
p
3/2
Cf r2 , where Cf
is the local skin friction coefficient and r2 is the ratio between the bulk and
free-stream temperatures (r2 ≈ 1 at subsonic speeds). Gordeyev et al. (2011a)
employed the linking equation along with the assumption of negligible pressure
fluctuations and SRA to develop a model for OP Drms at both subsonic and supersonic speeds. Their model takes the form OPDrms ∼ ρδ
p
Cf F1 (M ), where
F1 (M ) is a function of the mean and fluctuating velocity profiles and the freestream Mach number M ; F1 (M ) ≈ M 2 at subsonic speeds. This model was
shown to correctly predict both the amplitude and the convective velocity of
optically-active structures in compressible boundary layers. Both models (Wyckham & Smits 2009, Gordeyev et al. 2011a) were found to generally agree with
each other up to M = 5.
Non-adiabatic effects with heated and cooled walls were studied by Cress (2010)
and Cress et al. (2010). It was predicted and experimentally verified that proper
cooling of the wall upstream of the aperture will significantly decrease aero-optical
distortions. Large intermittent increases of aero-optical aberrations with subsequent drop-outs in far-field intensities have also been investigated experimentally
(Gordeyev et al. 2003, Cress 2010).
All these experimental studies strongly suggest that the majority of opticallyactive structures in compressible boundary layers are located in the outer re-
26
WANG, MANI & GORDEYEV
gion of the boundary layer, moving at 0.82–0.85 times the freestream velocity.
The predominant mechanism for density fluctuations in compressible boundary
layers is believed to be adiabatic heating/cooling due to velocity fluctuations
via strong Reynolds analogy, as pressure fluctuations inside boundary layers are
much smaller than temperature fluctuations. This was also confirmed in numerical simulations (Wang & Wang 2011).
While experimental investigations have shed much light on the characteristics and scaling laws of boundary-layer aero-optics, numerical simulations have
started to play a significant role in gaining physical understanding of distortion
mechanisms and testing aero-optical theories and their underlying assumptions.
High-fidelity approaches like DNS and well-resolved LES are ideally suited for
such fundamental studies because they provide detailed spatial and temporal information about the index-of-refraction field along with the velocity field, making
it possible to directly relate flow structures to optical aberrations.
Tromeur et al. (2003, 2006a, 2006b) undertook the first numerical investigation
of the optical-aberrating effects of turbulent boundary layers using LES. They
considered temporally and spatially evolving flat-plate boundary layers at freestream Mach numbers of 0.9 and 2.3 and momentum-thickness Reynolds number
Reθ = 2917. Converged flow and optical statistics were obtained and analyzed in
the case of spatially developing boundary layers with an adiabatic wall. By comparing directly computed OPDrms with that obtained using the linking equation,
Equation 8, with LES density-fluctuation data, Tromeur et al. (2006a, 2006b)
found significant discrepancies between the two, and consequently questioned the
applicability of the linking equation to boundary-layer flows. However, it was
later noted by Wang & Wang (2011) that these discrepancies were caused by
Aero-Optics
27
improper choice of the correlation-length definition that is inconsistent with the
linking equation.
Wang & Wang (2009, 2011) performed a well-resolved LES study of opticaldistortion mechanisms in a Mach 0.5, adiabatic-wall boundary layers at Reynolds
numbers Reθ = 875, 1770 and 3550. They obtained detailed statistics of fluctuating density and wavefront distortions, including the rms values, spatial and temporal correlations, and frequency spectra. Contributions from different boundarylayer regions to wavefront errors were evaluated, which shows dominance of the
logarithmic layer and wake region. Consistent with the analysis of Mani et al.
(2008), the effect of small-scale flow structures on optical aberrations was found to
be small. The applicability of Sutton’s linking equation was re-examined, and it
was found that, with a definition of the density correlation-length consistent with
the linking equation, the latter provides an excellent prediction of OPDrms . The
effect of turbulence inhomogeneity is too small to affect the validity of the linking
equation. The convection velocities of the optical wavefront were computed and
found to be consistent with previous experimental values.
It is well-established that turbulent boundary layers contain packets of vortical
structures with a preferred angular direction (e.g. Adrian 2007, Wu & Moin 2009).
They are believed to be important sources of aero-optical distortions, and therefore the boundary layer exhibits an anisotropic behavior for different elevation
angles of beam propagation. This behavior was measured experimentally (Cress
et al. 2008) and later observed computationally (Wang & Wang 2009, 2011). As
illustrated in Figure 3, a beam is distorted more severly when it is tilted toward
downstream than upstream with the same relative angle to the normal direction.
This is consistent with the linking-equation prediction, Equation 8, because the
28
WANG, MANI & GORDEYEV
correlation length Λ is longer when the beam is aligned with the elongated vortical
structures in the boundary layer (see right plot in Figure 3). Similar directional
dependence of optical distortions was observed by Truman & Lee (1990) in an
incompressible DNS of turbulent shear flow with uniform mean-shear rate.
Although it is generally believed that OPDrms is insensitive to Reynolds number at sufficiently high Reynolds numbers, the results of Wang & Wang (2011)
still exhibit a significant variation of OPDrms with Reynolds number (see Figure 3), although the rate of variation is seen to decrease with increasing Reynolds
number. Simulations at higher Reynolds numbers are required to allow direct,
quantitative comparisons with experimental measurements, which are typically
performed at much high Reynolds numbers.
5.2
Turbulent Shear Layers and Wakes
Separated turbulent shear layers are omnipresent around aero-optical devices
such as turrets and laser cavities. The presence of large-scale vortical structures
in shear layers was firmly established for the last 30 years or so, yet until few
years ago, most models used for aero-optical aberrations in shear layers largely
ignored the role of these structures, assuming that pressure fluctuations inside the
shear layer are negligible and using strong Reynold analogy to compute density
fluctuations. The reader is referred to Fitzgerald & Jumper (2004) for a detailed
survey of different shear layer models.
Vortical structures are characterized by a strong radial pressure gradient accompanied by significant pressure fluctuations. Based on this fact, Fitzgerald &
Jumper (2004) developed a physics-based numerical procedure for approximating
the density field, and thus the refractive index, from a known two-dimensional
Aero-Optics
29
velocity-field data for total-temperature-matched shear layers. The velocity, density and pressure fields are connected by the Euler equations, and by assuming
an isentropic process and using a simplified version of the energy equation and
the ideal gas law, they developed a procedure for iteratively retriving density and
pressure fileds based on the given velocity field. Their model, called the Weakly
Compressible Model (WCM), takes into account the radial pressure gradients required to sustain the vortical flow and curved path-lines. Fitzgerald & Jumper
(2004) showed that this model predicted reasonably well the experimentallyobserved aero-optical aberrations for a high subsonic shear layer, when other
shear-layer models failed to do so. However, as a simplified model, the accuracy
of WCM is limited by its underlying assumptions. A comparison with the numerical results of a compressible shear layer obtained by Visbal (2009) using ILES
with a high-order scheme shows that WCM, while correctly predicting density
fluctuations inside vortical structures, is unable to predict the sharp densitygradients numerically observed in braid regions between vortical structures.
To further investigate the nature of this discrepancy, simulations for laminar
mixing layers with either matched total or static temperature of the upper and
lower streams were performed by Visbal (2009). It was found that when the
total temperature was matched, a density interface between the two streams
was apparent in the shear layer braid regions, although this effect has not yet
been observed experimentally. In contrast, sharp density gradients were absent
when the static temperature was matched. Despite the significant differences
in the density field, the OPD distributions were found to be similar, indicating
that optical distortions are dominated by compressible effects instead of density
interfaces.
30
WANG, MANI & GORDEYEV
For transitional mixing layers, Visbal (2009) showed that optical distortions
near a splitter plate are similar to those caused by laminar mixing layers and
increase downstream as the mixing layer grows. The effects of forcing with either
spanwise uniform or non-uniform excitations on turbulent mixing layers have been
studied as well (Visbal 2008, White et al. 2010). It was shown that in general,
forced mixing layers induce larger OPD, and the forcing regularizes the shear
layer, confirming the experimental results of Rennie et al. (2008).
Mani et al. (2009) performed a fundamental study of optical distortions by
separated shear layers and wakes using a highly accurate LES with high grid
resolution. They considered a Mach 0.4 flow over a circular cylinder at Reynolds
numbers Re = 3900 and 10, 000. They found that optical distortions by the fully
developed wake was insensitive to the Reynolds number while distortions by the
separated shear layers were sensitive to the Reynolds number. This finding was
explained on the basis that the instability of the wake is predominantly inviscid
and that the Reynolds number only affects the Kolmogorov scale which was
found to be optically unimportant (Mani et al. 2008). In contrast, the instability
and transition in the early separated shear layers are Reynolds-number sensitive,
leading to Reynolds-number sensitivity of their optical effects.
The aero-optics of shear layers with different gases on both sides were studied
by Dimotakis et al. (2001), Catrakis & Aguirre (2004), and Zubair & Catrakis
(2007), to name a few. Dimotakis et al. (2001) studied subsonic shear layers at
low and high Reynolds numbers with different gases on the two sides. Gases were
density-matched, but had different index-of-refraction coefficients. Instantaneous
snapshots of index-of-refraction fields were taken using the Rayleigh scattering
technique and instantaneous wavefronts were calculated by integrating the field in
Aero-Optics
31
the beam-propagation direction. The large-scale structures in the shear layer were
found to be the dominant source for aero-optical distortions, and a level-set model
was proposed to explain observed wavefront characteristics. As pointed out by
Jumper & Fitzgerald (2001), the combined aero-optical effects were from both
index-of-refraction mixing and compressible turbulent mixing and, ultimately,
the proposed level-set model underestimated the experimentally-measured aerooptical levels of subsonic index-of-refraction-matched compressible shear layers
(Fitzgerald & Jumper 2004). Catrakis & Aguirre also used the Rayleigh scattering technique to study the aero-optical structure in dissimilar-gas mixing layers
at high Reynolds numbers. They developed a new interfacial fluid thickness approach, based on tracing regions of high gradients of the index-of-refraction field,
to study the physics of aero-optical distortions.
5.3
Flow over Optical Turrets
One of the most used platforms to point-and-track lasers are turrets of hemisphereon-cylinder type. They provide a convenient way to steer-and-stay the beam in
a desired direction for a transmitting station or to keep lock on the incoming
beam for a receiving station. Most airborne laser-based systems, both past and
present, use the turret geometry of some sort. But their less-than-ideal aerodynamic shape creates complex flow fields consisting of all major fundamental
turbulent flows: boundary layers, separated shear layers, wake, necklace vortex
and other large-scale vortical structures, as shown schematically in Figure 4.
Depending on the viewing direction, the laser beam will encounter one or several of these fundamental flows. Even the time-averaged flow around the turret
has significant steady pressure and density gradients, causing steady-lensing ef-
32
WANG, MANI & GORDEYEV
fects. Flow is typically attached on the front portion of the turret with a relatively
thin boundary layer, so the outgoing laser beam has mostly steady-lensing aberrations imposed on it. Flow is separated on the aft portion of the turret and,
when transmitting through the separated region, the beam will experience significant unsteady aero-optical aberrations even at relatively small subsonic speeds
(Gordeyev et al. 2007a, Gordeyev et al. 2007b, Vukasinovic et al. 2010, Gordeyev
et al. 2010b). Aero-optical distortions in this region are predominantly caused by
shear-layer vortical structures and the separation bubble formed downstream of
the turret (see Figure 4). At transonic and supersonic flows, shock-induced aerooptical effects with large density gradients are added to this already complicated
picture (Gordeyev & Jumper 2010).
Due to technological importance of turrets, they have been studied extensively
in the 1970s and 1980s, and good reviews of research efforts during that period
are presented in Gilbert & Otten (1982) and Sutton (1985). But laser systems of
those days had long wavelengths (∼ 10 microns), and unsteady aero-optical effects
caused by turrets were found to be negligible at subsonic speeds; only steadylensing effects were considered. The switch to shorter wavelength (∼ 1 micron)
laser in recent years has exacerbated the aero-optical effects, which spurred a
renewed interest in aero-optics of turrets. A summary of recent efforts, including flow-control strategies to mitigate aero-optical effects around turrets, can be
found in Gordeyev & Jumper (2010).
Recent experiments (Gordeyev & Jumper 2010) have shown that for subsonic
flows at large Reynolds numbers, OPDrms ∼ A(geometry, a/D)B(α, γ)ρ∞ M 2 D,
where ρ∞ and M are the free-stream density and Mach number, respectively,
D is the turret radius, A is a “constant” depending on the turret geometry and
Aero-Optics
33
the relative aperture size, a/D, and B is a function of the beam-transmitting
direction, defined by window angle α and elevation angle γ shown in Figure 4.
This simple scaling was found to collapse most of the experimental results available in open literature, as shown in Figure 5. The turret geometry was found
to influence aero-optical aberrations. A hemispherical turret is more opticallyaberrating than a hemisphere-on-cylinder turret due to increased proximity of
the near-the-wall vortices to the aperture. A flat-window turret creates more
aero-optical distortions compared to a conformal-window turret due to an earlier
separation from the front portion of the flat-window aperture.
Numerical simulations of flow over turrets are extremely challenging due to
the complex vortex structures discussed above and the wide range of flow scales
associated with them. The high Reynolds numbers encountered at realistic flight
or laboratory conditions make these flows out of reach by LES. In addition, simulations must cope with traditional difficulties such as the laminar-to-turbulence
transition of the boundary layer and incipient separation from the turret surface. Aside from a few LES attempts (Jones & Bender 2001, Arunajatesan &
Sinha 2005), which are grossly under-resolved, most recent computational efforts
are based on RANS/LES hybrid approaches. Nahrstedt et al. (2008) employed
the partially-averaged Navier-Stokes (PANS) method (Girimaji & Abdol-Hamid
2005) with the k-ε model in a finite-volume CFD code to compute the aerooptical flow and OPDrms of a 12-inch-diameter hemispherical/cylindrical turret
in a wind-tunnel experiment (Gordeyev et al. 2007b) at Mach numbers ranging
from 0.3 to 0.5. With a 2.7 million mesh, their results captured some aspects of
the experimental measurements, but sizable discrepancies existed, and the vortical structures in the shear and wake were apparently too coherent. Ladd et al.
34
WANG, MANI & GORDEYEV
(2009) used DES with the two-equation k-ω SST model to compute the same flow
with 3.2 million structured mesh and 10 million unstructured mesh, and obtained
improved solutions. Their mean turret-surface pressure and velocity profiles at
several wake stations showed reasonable agreement with experimental values, as
did their OPDrms over a range of elevation angles. The best resolved turret-flow
simulation to date was performed by Morgan & Visbal (2010), who again considered the experimental configuration of Gordeyev et al. (2007b). Using a hybrid
RANS/ILES method with a k-ε model and a 6th-order compact-difference solver
on a 23.6 million mesh, their simulation captured fine details of turbulence structures and many flow features observed in the experiment, but there were also
significant differences in the separated region. Major discrepancies were observed
between the computed OPDrms values and experimental data.
Virtually all numerical simulations of turret flows to date are for turrets with a
conformal window. Flows around flat-window turrets are more difficult to study
both experimentally and computationally. The slope-discontinuity around the
flat aperture causes a premature flow separation and creates a weak separation
bubble over the aperture at some angles (Gordeyev et al. 2007a, Gordeyev et
al. 2010b), leading to increased aero-optical distortions. Aside from increased
geometric and flow complexity, computational expenses are higher due to the
fact that a separate flow simulation is required for each viewing angle because
the flow field around the turret depends on the window position.
Evidently, current computational capabilities still do not allow accurate and
robust predictions of the fluctuating index-of-refraction field around a threedimensional optical turret at realistic Reynolds numbers. Hybrid LES/RANS
methods continue to offer the best practical approach in the foreseeable future,
Aero-Optics
35
but a number of weaknesses and outstanding issues need to be addressed. For
DES methods, which treat attached boundary layers with RANS, prediction of
boundary-layer transition and incipient separation from the turret surface is a
significant challenge. Furthermore, the shear layer immediately after separation,
which is in the optically important region, may not behave realistically without
upstream turbulence. The effect of upstream turbulent fluctuations on the shearlayer instability and how to incorporate them in DES-type simulations requires
further investigation.
Given the significant computational challenges for three-dimensional turrets, it
would be prudent to first validate solution techniques in simpler flows that capture some crucial physical aspects of realistic turret flows. The two-dimensional
cylindrical turret investigated experimentally by Gordeyev et al. (2011b) provides
an excellent benchmark configuration for computational validation. Because the
turret is mounted horizontally on two flat surfaces of different altitude, the flow
upstream of the turret is a flat-plate turbulent boundary layer, which can be
easily characterized and realistic turbulent inflow can be fed into the simulation
(Wang & Wang 2009). This configuration has been employed in a number of recent LES investigations (Morgan & Visbal 2010,Wang & Wang 2009,Wang et al.
2010).
6
CONCLUDING REMARKS
Aero-optics is an interdisciplinary area that has seen a renewed interest and significant growth over the past decade. At the core of aero-optical phenomena is
compressible turbulence, whose accurate prediction in a practical aero-optical environment over optically active flow scales remains a significant challenge. This
36
WANG, MANI & GORDEYEV
review is focused on the fundamental understanding and prediction of aero-optics,
and related numerical and experimental techniques. Given the diverse nature of
the field and space limitation, the contents in this article are necessarily selective;
they are by no means reflective of the relative level of research activities or accomplishments. Computational techniques and their applications to aero-optical
problems are afforded more coverage considering the growing role of computations
and their future promise.
The discussions presented in this paper are mostly on the aero-optical effects
of subsonic flows. The presence of shock waves and shock-turbulence interaction
in a supersonic flow exacerbates aero-optical effects and gives rise to additional
computational and experimental challenges, which are not discussed here. Wavefront sensors and data reduction algorithms are only briefly mentioned, but they
are the most critical part in collecting accurate aero-optical measurements.
Another important part of aero-optics research not covered in the present article is mitigation strategies. The goal of aero-optical mitigations is to reduce the
overall distortions of an optical beam, which can be achieved through a variety of
flow control and/or adaptive optic techniques. Passive control devices placed in
front of an optical turret have shown promise (Gordeyev et al. 2010a), and active
flow control using pulsating micro-jets has effectively reduced aero-optical distortions around turrets (Vukasinovic et al. 2010). Although adaptive optics alone
currently only works for low-frequency aberrations typical of atmospheric turbulence, progress has been made in extending it to high-frequency aero-optics. One
promising technique, which has been demonstrated in a compressible shear layer
(Rennie at al. 2008), involves a combination of adaptive optics with flow control.
By first regularizing the flow using active flow contol, the aberrations become
Aero-Optics
37
predictable and can be corrected by a deformable mirror without instantaneous
wavefront measurements.
In the development of mitigation techniques for optical systems, fundamental
understanding of aero-optical mechanisms gained through physical and numerical experiments has played and will continue to play an essential role. Further
advances in experimental and computational techniques leading to acccurate predictive tools are much needed in order to accelerate the development of practical
airborne laser systems.
ACKNOWLEDGMENTS
The authors acknowledge support from the Air Force Office of Scientific Research,
the High Energy Laser Joint Technology Office, and the Air Force Research Laboratory for research in this area. We would like to thank Eric Jumper and colleagues in the Aero-Optics Group at the University of Notre Dame and Parviz
Moin of Stanford University for many fruitful discussions.
Literature Cited
Abado S, Gordeyev S, Jumper E. 2010. Two-dimensional high-bandwidth Shack-Hartmann
wavefront sensor: design guidelines and evaluation testing. Opt. Eng. 49(6):064403.
Adrian RJ. 2007. Hairpin vortex organization in wall turbulence. Phys. Fluids 19(4):1-16.
Arunajatesan S, Sinha N. 2005. Analysis of line of sight effects in distortions of laser beams
propagating through a turbulent turret flow field. AIAA Pap. 2005-1081
Blanchard P, Fisher D, Woods SC, Greenaway AH. 2000. Phase-diversity wave-front sensing
with a distorted diffraction grating. Appl. Optics, 39(35):6649–55
Born M, Wolf, E. 2002. Principles of Optics, 7th ed. Cambridge: Cambridge Univ. Press. 952
38
WANG, MANI & GORDEYEV
pp.
Cassady, PE, Birch, SF, Terry, PJ. 1989. Aero-optical analysis of compressible flow over an
open cavity. AIAA J. 27(6):758–62
Catrakis HJ, Aguirre RC. 2004. New interfacial fluid thickness approach in aero-optics with
applications to compressible turbulence. AIAA J. 43(10):1973–81
Chernov LA. 1960. Wave Propagation in a Random Medium. New York: McGraw-Hill. 168 pp.
Childs RE. 1993 Prediction and control of turbulent aero-optical distortion using large eddy
simulation. AIAA Pap. 93-2670
Coles WmA, Filice JP, Frehlich RG, Yadlowsky M. 1995. Simulation of wave propagation in
three-dimensional random media. Appl. Optics 34(12):2089–101
Cress J, Gordeyev S, Post M, Jumper EJ. 2008. Aero-optical measurements in a turbulent,
subsonic boundary layer at different elevation angles. AIAA Pap. 2008-4214
Cress JA. 2010. Optical aberrations caused by coherent structures in a subsonic, compressible,
turbulent boundary layer. Ph.D. Thesis, Univ. of Notre Dame, Notre Dame, IN
Cress J, Gordeyev S, Jumper E. 2010. Aero-optical measurements in a heated, subsonic, turbulent boundary layer. AIAA Pap. 2010-0434
Dimotakis PE, Catrakis HJ, Fourguette DC. 2001. Flow structure and optical beam propagation
in high-Reynolds number gas-phase shear layer and jets. J. Fluid Mech. 433:105–34
Fitzgerald EJ, Jumper EJ. 2002. Scaling aerooptic aberrations produced by high-subsonic-Mach
shear layers. AIAA J. 40(7):1373–81
Fitzgerald EJ, Jumper EJ. 2004. The optical distortion mechanism in a nearly incompressible
free shear layer. J. Fluid Mech. 512:153–89
Frehlich R. 2000. Simulation of laser propagation in a turbulent atmosphere. Appl. Optics
39(3):393–97
Germano M, Piomelli U, Moin P, Cabot WH. 1991. A dynamic subgrid-scale eddy viscosity
model. Phys. Fluids A 3(7):1760–65
Gilbert KG, Otten LJ. eds. 1982. Aero-Optical Phenomena, Progress in Astronautics and Aeronautics, Vol. 80. New York: AIAA. 412 pp.
Girimaji SS, Abdol-Hamid KS. 2005 Partially-averaged Navier-Stokes model for turbulence:
implementation and validation. AIAA Pap. 2005-502
Aero-Optics
39
Goodman JW. 2004. Introduction to Fourier Optics, 3rd ed. Roberts & Company. 466 pp.
Gordeyev S, Jumper E, Ng T, Cain A. 2003. Aero-optical characteristics of compressible, subsonic turbulent boundary layer. AIAA Pap. 2003-3606
Gordeyev S, Hayden T, Jumper E. 2007a Aero-optical and flow measurements over a flatwindowed turret. AIAA J. 45(2):347–57
Gordeyev S, Post M, MacLaughlin J, Ceniceros J, Jumper E. 2007b Aero-optical environment
around a conformal-window turret. AIAA J. 45(7):1514–24
Gordeyev S, Jumper EJ. 2010. Fluid dynamics and aero-optics of turrets. Prog. Aerospace Sci.
46:388–400
Gordeyev S, Cress JA, Smith A, Jumper EJ. 2010a. Improvement in optical environment over
turrets with flat window using passive flow control. AIAA Pap. 2010-4492
Gordeyev S, Jumper E, Vukasinovic B, Glezer A, Kibens V. 2010b. Hybrid flow control of a
turret wake, part II: aero-optical effects. AIAA Pap. 2010-0438
Gordeyev S, Jumper EJ, Hayden T. 2011a. Aero-Optics of Supersonic Boundary Layers. AIAA
Pap. 2011-1325
Gordeyev S, Cress JA, Jumper EJ, Cain AB. 2011b. Aero-optical environment around a cylindrical turret with a flat window. AIAA J. 49(2):308–15
Hardy JW. 1998. Adaptive Optics for Astronomical Telescopes. Oxford: Oxford Univ. Press.
438 pp.
Havener G. 1992. Optical wave front variance: a study on analytic models in use today. AIAA
Pap. 92-0654
Hugo, RJ, Jumper EJ, 1996. Experimental measurement of a time-varying optical path difference
by the small-aperture beam technique. Appl. Optics 35(22):4436-47
Hugo RJ, Jumper EJ. 2000. Applicability of the aero-optic linking equation to a highly coherent,
transitional shear layer. Appl. Optics 39(24):4392–401
Jones MI, Bender EE. 2001 CFD-based computer simulation of optical turbulence through
aircraft flowfields and wakes. AIAA Pap. 2001-2798
Jumper EJ, Fitzgerald EJ. 2001. Recent advances in aero-optics. Prog. Aerospace Sci. 37:299–
339
Ladd J, Mani M, Bower W. 2009 Validation of aerodynamic and optical computations for the
40
WANG, MANI & GORDEYEV
flow about a cylindrical/hemispherical turret. AIAA Pap. 2009-4118
Liepmann HW. 1952. Deflection and diffusion of a light ray passing through a boundary layer.
Tech. Rep. SM-14397, Douglas Aircraft Company, Santa Monica Division, Santa Monica, CA
Lilly DK. 1992. A proposed modification of the Germano subgrid scale closure method. Phys.
Fluids A 4(3):633–35
Lloyd-Hart M. 2003. Taking the twinkle out of starlight. IEEE Spectrum 40(12):22–29
Mahajan VN 1982. Strehl ratio for primary aberrations: some analytical results for circular and
annular pupils. J. Opt. Soc. Am. 72(9):1258–66
Mahajan VN 1983. Strehl ratio for primary aberrations in terms of their aberration variance J.
Opt. Soc. Am. 73(6):860–61
Mani A, Wang M, Moin P. 2006. Statistical description of free-space propagation of highly
aberrated optical beams. J. Opt. Soc. Am. A 23(12):3027–35
Mani A, Wang M, Moin P. 2008. Resolution requirements for aero-optical simulations. J.
Comput. Phys. 227:9008–20
Mani A, Moin P, Wang M. 2009. Computational study of optical distortions by separated shear
layers and turbulent wakes. J. Fluid Mech. 625:273–98
Maréchal A. 1947. Étude des effets combinés de la diffraction et des aberrations géométriques
sur l’image d’un point lumineux Rev. Opt. Theor. Instrum. 26:257–77
Moin P, Squires KD, Cabot WH, Lee S. 1991. A dynamic subgrid-scale model for compressible
turbulence and scalar transport. Phys. Fluids A 3(11):2746–57
Moin P, Mahesh K. 1998. Direct numerical simulation: a tool in turbulence research. Annu.
Rev. Fluid Mech. 30:539-78
Monin AS, Yaglom, AM. 1975. Statistical Fluid Mechanics: Mechanics of Turbulence, Vol. II,
ed. JL Lumley, Chap. 9, Cambridge, Mass: MIT Press. 886 pp.
Morgan PE, Visbal MR. 2010. Large eddy simulation of flow over a flat window cylindrical
turret with passive flow control. AIAA Pap. 2010-916
Morkivin MV. 1962. Effects of compressibility on turbulent flows. In Mechanique de la Turbulence, ed. A. Favre, CNRS, pp. 367–380. Paris, France
Nahrstedt D, Hsia Y-C, Jumper E, Gordeyev S, Ceniceros J, Weaver L, DeSandre L, McLaughlin
T. 2008. Wind tunnel validation of computational fluid dynamics-based aero-optics model.
Aero-Optics
41
Proc. IMechE Part G: J. Aerosp. Eng. 223(G4):393-406
Piomelli U, Balaras E. 2002. Wall-layer models for large-eddy simulations. Ann. Rev. Fluid
Mech. 34:349–74
Pond JE, Sutton GW. 2006 Aero-optic performance of an aircraft forward-facing optical turret.
J. Aircraft 43(3):600–07
Porter C, Gordeyev S, Zenk M, Jumper E. 2011. Flight measurements of aero-optical distortions
from a flat-windowed turret on the airborne aero-optics laboratory (AAOL). To be presented
at 42nd AIAA Plasmadynamics and Lasers Conference, Hawaii, June 2011.
Rennie RM, Duffin DA, Jumper EJ, 2008. Characterization and aero-optic correction of a forced
two-dimensional weakly compressible shear layer. AIAA J. 46(11):2787–95.
Rose WC. 1979. Measurements of aerodynamic parameters affecting optical performance. Air
Force Weapons Laboratory Final Report AFWL-TR-78-191, Kirtland AFB, NM
Ross TS. 2009. Limitations and applicability of the Maréchal approximation. Appl. Optics
48(10):1812–18
Sasiela RJ. 2007. Electromagnetic Wave Propagation in Turbulence: Evaluation and Application
of Mellin Transforms. 2nd ed. Bellingham, WA: SPIE Press. 369 pp.
Sinha N, Arunajatesan S, Seiner JM, Ukeiley LS. 2004. Large-eddy simulations of aero-optic
flow fields and control application. AIAA Pap. 2004-2448
Smagorinsky J. 1963. General circulation experiments with the primitive equations. I. The basic
experiment. Mon. Weath. Rev. 91:99–164
Smith RR, Truman CR, Masson BS. 1990. Prediction of optical phase degradation using a turbulent transport equation for the variance of index-of-refraction fluctuations. AIAA Pap. 900250
Spalart PR. 2009. Detached-eddy simulation. Annu. Rev. Fluid Mech. 41:181-202
Stine HA, Winovich W. 1956 Light diffusion through high-speed turbulent boundary layers.
Research Memorandum A56B21, NACA, Washington
Steinmetz WJ. 1982. Second moments of optical degradation due to a thin turbulent layer. In
Aero-Optical Phenomena, Progress in Astronautics and Aeronautics, Vol. 80, ed. KG Gilbert,
LJ Otten, pp. 78–100. New York: AIAA.
Sutton GW. 1969. Effects of turbulent fluctuations in an optically active fluid medium. AIAA
42
WANG, MANI & GORDEYEV
J. 7(9):1737-43
Sutton GW. 1985. Aero-optical foundations and applications. AIAA J. 23(10):1525–37
Tatarski VI. 1961. Wave Propagation in a Turbulent Medium. New York: McGraw-Hill. 285 pp.
Tromeur E, Garnier E, Sagaut P, Basdevant C. 2003. Large eddy simulations of aero-optical
effects in a turbulent boundary layer. J. Turbul. 4(005):1–22
Tromeur E, Garnier E, Sagaut P. 2006a. Large eddy simulation of aero-optical effects in a
spatially developing turbulent boundary layer. J. Turbul. 7(1):1–28
Tromeur E, Garnier E, Sagaut P. 2006b. Analysis of the Sutton model for aero-optical properties
of compressible boundary layers. J. Fluids Eng. 128:239–46
Truman CR, Lee MJ. 1990. Effects of organized turbulence structures on the phase distortion
in a coherent optical beam propagating through a turbulent shear flow. Phys. Fluids A
2(5):851–57
Truman CR. 1992. The influence of turbulent structure on optical phase distortion through
turbulent shear flows. AIAA Pap. 92-2817
Tsai YP, Christiansen WH. 1990. Two-dimensional numerical simulation of shear-layer optics.
AIAA J. 28(12):2092–97
Tyson RK. 1997. Principles of Adaptive Optics. 2th ed. Boston: Academic Press. 345 pp.
Tyson RK, ed. 2000. Adaptive Optics Engineering Handbook. New York: Marcel Dekker. 339
pp.
Visbal MR. 2008. Effect of flow excitation on aero-optical aberration. AIAA Pap. 2008-1074
Visbal MR. 2009. Numerical simulation of aero-optical aberration through weakly-compressible
shear layers. AIAA Pap. 2009-4298
Vukasinovic B, Glezer A, Gordeyev S, Jumper E, Kibens V. 2010 Fluidic control of a turret
wake: aerodynamic and aero-optical effects. AIAA J. 48(8):1686–99
Wang M, Moin P. 2002. Dynamic wall modeling for large-eddy simulation of complex turbulent
flows. Phys. Fluids 14(7):2043–51
Wang K, Wang M. 2009. Numerical simulation of aero-optical distortions by a turbulent boundary layer and separated shear layer. AIAA Pap. 2009-4223
Wang K, Wang M, Gordeyev S, Jumper E. 2010. Computation of aero-optical distortions over
a cylindrical turret with passive flow control. AIAA Pap. 2010-4498
Aero-Optics
43
Wang K, Wang M. 2011. Aero-optical distortions by subsonic turbulent boundary layers. Submitted
White MD. 2010. High-order parabolic beam approximation for aero-optics. J. Comput. Phys.
229:5465–85
White MD, Morgan PE, Visbal MR. 2010. High-fidelity aero-optical analysis. AIAA Pap. 2010433
White MD, Visbal MR. 2010. High-fidelity analysis of aero-optical interaction with compressible
boundary layers. AIAA Pap. 2010-4496
Wolters DJ. 1973. Aerodynamic effects on airborne optical systems. Tech. Rep. MDC A2582,
McDonnell Douglas Corporation, St. Louis, MO
Wyckham CM, Smits AJ. 2009. Aero-optic distortion in transonic and hypersonic turbulent
boundary layers. AIAA J. 47(9):2158–68
Wolfe WL, Zizzis GJ. 1978. The Infrared Handbook. Office of Naval Research, Department of
the Navy.
Wu X. 2009. Direct numerical simulation of turbulence in a nominally-zero-pressure-gradient
flat-plate boundary layer. J. Fluid Mech. 630:5–41
Zubair FR, Catrakis HJ. 2007 Aero-optical resolution robustness in turbulent separated shear
layers at large Reynolds numbers. AIAA J. 45(11):2721–28
44
WANG, MANI & GORDEYEV
Figure 1: Schematic of the aero-optics problem.
45
Aero-Optics
−6
−5
-10
y/D
σy
-10
y/D
10
λ/D=10
10
λ/D=2.5×10
σx
20
0
x/D
20
0
x/D
20
0
x/D
20
y/D
-10
-10
y/D
10
x/D
10
0
0.001
0.0008
0.0006
0.0004
0.0002
0
Figure 2: Instantaneous far-field optical irradiance for Gaussian beams through
the wake of a circular cylinder of diameter D at a distance of 105 D for two
optical wavelengths. The top plots show the distorted beams, and the bottom
plots show the undistorted, diffraction limited beams. The aperture diameter
is a = 0.3D. The peak dimensionless intensity (relative to the aperture value)
for the undistorted beams are 0.074 and 0.005 for λ/D = 2.5 × 10−6 and 10−5 ,
respectively. The results are obtained using LES at free-stream Mach number
M = 0.4 and Reynolds number ReD = 3900 (Mani et al. 2009).
46
WANG, MANI & GORDEYEV
11
y + = 198
OPDrms /δ × 107
10
9
∆y/δ
Reθ = 3550
Reθ = 1770
Reθ = 875
0.9
0.7
0.5
0.3
0.1
0.2
0
8
-0.5
0
∆x/δ
0.5
7
0.4
∆y/δ
6
5
4
60
80
100
β
120
y + = 50
0.9
0.7
0.5
0.3
0.1
0.2
0
-0.5
0
∆x/δ
0.5
Figure 3: Left: OPDrms variation with elevation angle (measured from upstream
wall) for an optical beam passing through a Mach 0.5 turbulent boundary layer at
three different Reynolds numbers. The aperture size is 7δ in streamwise (x) and
2.4δ in spanwise (z) directions. Right: two-point spatial correlations of density
fluctuations in the x-y plane at two wall-normal locations for Reθ = 3550. The
results are obtained from LES (Wang & Wang 2011).
Aero-Optics
47
Figure 4: Flow topology around an optical turret and measurement coordinates,
from Gordeyev & Jumper (2010).
48
WANG, MANI & GORDEYEV
Figure 5: Normalized levels of aero-optical distortions induced by turret flows
measured for different viewing angles and turret geometries (Gordeyev & Jumper
2010).